License: arXiv.org perpetual non-exclusive license
arXiv:2209.14445v3 [cond-mat.mes-hall] 29 Dec 2023

Electrical manipulation of valley-qubit and valley geometric phase
in lateral monolayer heterostructures

Jarosław Pawłowski [email protected] Department of Theoretical Physics, Wrocław University of Science and Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland    John Eric Tiessen Department of Electrical and Computer Engineering, University of Illinois at Chicago, Chicago, IL 60607, USA    Rockwell Dax Department of Electrical and Computer Engineering, University of Illinois at Chicago, Chicago, IL 60607, USA    Junxia Shi Department of Electrical and Computer Engineering, University of Illinois at Chicago, Chicago, IL 60607, USA
Abstract

We explore a solid state qubit defined on valley isospin of an electron confined in a gate-defined quantum dot created in an area of monolayer MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT/WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT lateral junction, where a steep dipolar potential emerges. We show that the junction oriented along an armchair direction can induce intervalley transitions of the electron confined in the neighboring quantum dot when the (gate-controllable) overlapping with the junction is significant and pumping frequency tuned. The pumping scheme that induces transitions is all-electrical: obtained by applying oscillating voltages to control gates and thus enables for scalable qubit architectures. We also report another possibility of valley-qubit manipulation by accumulating non-Abelian valley Berry phase. To model nanodevice we solve the time-dependent Schrödinger-Poisson equations in a tight-binding approach and obtain exact time-evolution of the valley-qubit system.

I Introduction

Advances continue to be made in the implementation of quantum computing hardware. However, much of the recent progress in quantum computing chips has been made with superconducting qubits [1]. Such implementations of quantum computing hardware have so far been limited to dozens of qubits rather than the thousands needed for fault tolerant quantum computing architectures [2]. Therefore, alternative types of qubits, as well as different qubit host materials, have been an area of great interest in the last several years [3, 4, 5]. Most of these proposed qubit systems seek to use silicon as the host material since it is well understood, mainly due to its long history of use in the semiconductor industry. However, silicon has intrinsic limitations, such as relatively weak spin-orbit coupling, or isotopic impurities leading to spin dechoherence, which also restrict its utility as a qubit host. These issues can be remedied [6, 7], nonetheless, a more suitable qubit host material should be found.

2D materials such as Transition Metal Dichalcogenides (TMDCs) and Bi-Layer Graphene (BLG) offer an alternative approach to the problem of manipulating and isolating qubits. TMDCs in particular are an excellent candidate for qubit hosting and manipulation [8, 9, 10, 11, 12, 13, 14, 15] due to their intrinsic spin-orbit interaction and 2D nature. In this work we propose a 2D lateral TMDC heterostructure [16] to act as a qubit host material, while using the position dependent change in the electronic properties of the heterostructure near the interface to control the state of the qubit. Due to recent progress in the growth of lateral TMDC heterostructures [17, 18], such a structure seems reasonable from an experimental/engineering perspective [16, 19]. Also, the all-electric nature of the proposed qubit means that the footprint needed for device implementation can be brought down to the order of nanometers.

Examples of 2D lateral TMDC heterostructures include the MoS(Se)22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT/WS(Se)22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT semiconductor heterostructures [20, 21, 22, 23, 24, 25]. In particular, MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT/WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT [21] and MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT/WSe22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT [23, 25] heterostructures with an atomically sharp interface in the armchair or zigzag configuration have been successfully synthesized. Also interesting physical properties for device applications using TMDC lateral junctions, such as the photovoltaic effect, have been demonstrated experimentally [17, 16]. Such structures also open the way to develop ultra-efficient, planar thermoelectric devices [26]. An important limitation, however, is that lateral interfaces can only be realized in epitaxially grown TMDCs, which are known to exhibit lower quality than mechanically exfoliated crystals.

In addition to the spin degree of freedom, TMDCs also possess a valley isospin. This additional degree provides an extra set of states for defining qubits, but the problem arises of how to efficiently manipulate such qubit states. To achieve this, one may use optical manipulation [27], but for a scalable architecture electric control is desired. To control the qubit via local gating, creation of a confinement potential with a size comparable to the lattice vector is needed [28, 29, 11, 14], which is difficult to engineer using presently available gate lithography resolutions.

In our proposed system, this problem is overcome via the utilization of the sharp potential profile naturally generated at the lateral heterojunction of the two 2D materials. Such an inter-TMDC-monolayer interface (assuming that it is atomically sharp and of armchair type) generates a steep dipolar potential that changes the momentum profile of the gate-induced confining potential for the electron, which enables the coupling between the states in K𝐾Kitalic_K and Ksuperscript𝐾K^{\prime}italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT valleys. The magnitude of this effect depends on the distance between the electron and the junction that can be varied by changing the applied gate voltage.

The paper is organized as follows: In Section II we will introduce the device structure based on the TMDC heterojunction, as well as the theoretical model which describes spin-valley physics of a single-electron qubit carrier confined within the junction area. In the next Section III we present results on valley qubit manipulation, while in Section IV we analyze conditions that should be met to couple valley states. In the final Section, V, we study the possibility of valley manipulation by geometric Berry phase accumulation.

II Device model

Refer to caption
Figure 1: The schematic view on the proposed device structure containing MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT/WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT heterojunction (orange/blue) sandwiched between hBN tunnel barriers, together with four gates G{L,R,U,D}subscriptGLRUD\mathrm{G}_{\mathrm{\{L,R,U,D\}}}roman_G start_POSTSUBSCRIPT { roman_L , roman_R , roman_U , roman_D } end_POSTSUBSCRIPT deposited on top defining quantum dot in the junction area.

The proposed nanodevice is composed of a MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT/WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT lateral (in-plane) heterojunction sandwiched between hBN tunnel barriers each 5555 nm thick – see Fig. 1. On top of the upper hBN layer four metallic gates are deposited, G{L,R,U,D}subscriptGLRUD\mathrm{G}_{\mathrm{\{L,R,U,D\}}}roman_G start_POSTSUBSCRIPT { roman_L , roman_R , roman_U , roman_D } end_POSTSUBSCRIPT. Their role is to create a quantum dot (QD) confinement potential in the junction area, and to control valley coupling (via GLsubscriptGL\mathrm{G}_{\mathrm{L}}roman_G start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT and GRsubscriptGR\mathrm{G}_{\mathrm{R}}roman_G start_POSTSUBSCRIPT roman_R end_POSTSUBSCRIPT) by tuning the confined electron’s wavefunction overlap with the junction dipole potential profile. The whole structure is placed on the highly doped substrate which acts as a backgate with zero reference voltage. Profiles of the QD potential defined by the top gates, and calculated using the Poisson-Schrödinger method, are presented in Fig. 2(a) – cross-section at four top gates level, zg=10subscript𝑧𝑔10z_{g}=10italic_z start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 10 nm, and Fig. 2(b) – cross-section at monolayers level, zl=5subscript𝑧𝑙5z_{l}=5italic_z start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = 5 nm with also visible junction potential profile at xj=7subscript𝑥𝑗7x_{j}=7italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 7 nm. In Fig. 2(c) with presented lateral x𝑥xitalic_x-z𝑧zitalic_z cross-section through the device we can observe electric dipole at the junction (zoomed inset) and constant potential conditions by applying voltages to the GLsubscriptGL\mathrm{G}_{\mathrm{L}}roman_G start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT and GRsubscriptGR\mathrm{G}_{\mathrm{R}}roman_G start_POSTSUBSCRIPT roman_R end_POSTSUBSCRIPT top gates. Also top gates GUsubscriptGU\mathrm{G}_{\mathrm{U}}roman_G start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT and GDsubscriptGD\mathrm{G}_{\mathrm{D}}roman_G start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT are visible but in cross-section in y𝑦yitalic_y-z𝑧zitalic_z directions presented in Fig. 2(d). It should also be added that in realistic configurations, the thickness of the hBN spacer can be increased, e.g. to 10-15 nm, to avoid voltage breakdown, but this may slightly reduce the electron confinement.

Refer to caption
Figure 2: Electrostatic potential profiles within the nanodevice. Cross-sections: (a) at the top gates level (with marked gates positions), (b) monolayers level with visible junction profile, or along (c) y=0𝑦0y=0italic_y = 0 surface with visible junction dipole (inset), and (d) x=0𝑥0x=0italic_x = 0 surface.

II.1 Heterojunction

To realistically model the junction between two 2D monolayer materials we have made the following assumptions. An MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT/WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT junction possess type-II band alignment [30] with a conduction band (CB) offset of approximately 0.35similar-toabsent0.35\sim 0.35∼ 0.35 eV [31, 32]. This can be calculated using the Anderson rule and the difference in electron affinity of these two materials. In addition to the band offset caused by connecting the two semiconductors, the Fermi levels of the two semiconductors must also match. This matching leads to characteristic band bending in the junction area and determines the exact shape of the junction potential profile. Fermi level misalignment generates charge flow until EFsubscript𝐸𝐹E_{F}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT equilibrates creating space charges (charge transfer across the interface [33, 34] from MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT to WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT) that build up forming the electrostatic dipole at the junction [32]. This induced dipole is strongly localized to the heterojunction. Therefore, to model this effect we assume two charge densities which form a dipole linear charge density with a value of λ=378|e|/nm𝜆378𝑒nm\lambda=378\,|e|/\mathrm{nm}italic_λ = 378 | italic_e | / roman_nm. Next, we assume the following density distribution for space charge:

d±(x,y,z)=±λ2πσxσze(xxj±δ)22σx2(zzl)22σz2,superscript𝑑plus-or-minus𝑥𝑦𝑧plus-or-minus𝜆2𝜋subscript𝜎𝑥subscript𝜎𝑧superscript𝑒superscriptplus-or-minus𝑥subscript𝑥𝑗𝛿22superscriptsubscript𝜎𝑥2superscript𝑧subscript𝑧𝑙22superscriptsubscript𝜎𝑧2d^{\pm}(x,y,z)=\frac{\pm\lambda}{2\pi\sigma_{x}\sigma_{z}}e^{-\frac{(x-x_{j}% \pm\delta)^{2}}{2\sigma_{x}^{2}}-\frac{(z-z_{l})^{2}}{2\sigma_{z}^{2}}},italic_d start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT ( italic_x , italic_y , italic_z ) = divide start_ARG ± italic_λ end_ARG start_ARG 2 italic_π italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT - divide start_ARG ( italic_x - italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ± italic_δ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG ( italic_z - italic_z start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT , (1)

with σx=σz=0.1subscript𝜎𝑥subscript𝜎𝑧0.1\sigma_{x}=\sigma_{z}=0.1italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0.1 nm, and charge displacement δ=0.25𝛿0.25\delta=0.25italic_δ = 0.25 nm. The linear density λ𝜆\lambdaitalic_λ parameter was tuned in our model to give the desired CB offset 0.350.350.350.35 eV. Its actual value is about 3 times larger than the one derived from [35]: λ=130|e|/nm𝜆130𝑒nm\lambda=130\,|e|/\mathrm{nm}italic_λ = 130 | italic_e | / roman_nm. We take d+(x,y,z)superscript𝑑𝑥𝑦𝑧d^{+}(x,y,z)italic_d start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_x , italic_y , italic_z ) at MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT side, d(x,y,z)superscript𝑑𝑥𝑦𝑧d^{-}(x,y,z)italic_d start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_x , italic_y , italic_z ) at WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT side and 0 elsewhere in computational box (i.e. outside the monolayers).

The resulting spatial charge dipole density is presented in Fig. 3(b). This dipole generates a built-in electric field across the interface as seen in Fig. 3(a) (solid lines). The obtained CB offset is of the desired value. Also it should be noted that we assume that the difference in electron affinities and work functions are similar and therefore the potential levels on both sides far from the junction match with each other.

Refer to caption
Figure 3: (a) Potential profiles (solid lines) through the junction interface, with the characteristic CB offset built by space charge dipole (b). Modulation of top gate voltages changes the potential, also in direction along the junction (c), and shifts electron (dashed lines) confined within the QD.

If we additionally modulate the voltage applied to the left gate GLsubscriptGL\mathrm{G}_{\mathrm{L}}roman_G start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT we may tune the shape of the confinement potential (mostly at the MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT side) which controls the position of the electron density (dashed lines) confined in the QD. In this way by changing VLsubscript𝑉LV_{\mathrm{L}}italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT we may control the amount of the electron density which is localized at the junction. A similar control scheme may by done in the y𝑦yitalic_y direction by instead using the voltages VUsubscript𝑉UV_{\mathrm{U}}italic_V start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT and VDsubscript𝑉DV_{\mathrm{D}}italic_V start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT – see Fig. 3(c).

Sulphur vacancies are found to be the dominant defect in MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT and WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT[32]. The implemented junction model assumes pristine monolayer materials, however in the case of defects/dopants present in monolayers which change at the Fermi levels, additional source-drain bias will be needed to restore the desired junction profile.

II.2 Tight-binding model

To model a single-electron confined in the in-plane monolayer junction area, we develop an atomistic tight-binding (TB) model with different parameters for both sides of the heterostructure and average the hopping parameter through the interface [36, 16]:

H=𝐻absent\displaystyle H=italic_H = mA(B)ασ(ϵαA(B)+φm)c^mασc^mασsubscript𝑚𝐴𝐵𝛼𝜎subscriptsuperscriptitalic-ϵ𝐴𝐵𝛼subscript𝜑𝑚superscriptsubscript^𝑐𝑚𝛼𝜎subscript^𝑐𝑚𝛼𝜎\displaystyle\,\sum_{\begin{subarray}{c}m\in A(B)\\ \alpha\,\sigma\end{subarray}}(\epsilon^{A(B)}_{\alpha}+{\varphi}_{m})\,\hat{c}% _{m\alpha\sigma}^{\dagger}\hat{c}_{m\alpha\sigma}∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_m ∈ italic_A ( italic_B ) end_CELL end_ROW start_ROW start_CELL italic_α italic_σ end_CELL end_ROW end_ARG end_POSTSUBSCRIPT ( italic_ϵ start_POSTSUPERSCRIPT italic_A ( italic_B ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + italic_φ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_m italic_α italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_m italic_α italic_σ end_POSTSUBSCRIPT (4)
+\displaystyle++ mA(B)αβσσsσσzλA(B)αβc^mασc^mβσ+HZ+HRsubscript𝑚𝐴𝐵𝛼𝛽𝜎superscript𝜎subscriptsuperscript𝑠𝑧𝜎superscript𝜎subscriptsuperscript𝜆𝐴𝐵𝛼𝛽subscriptsuperscript^𝑐𝑚𝛼𝜎subscript^𝑐𝑚𝛽superscript𝜎subscript𝐻𝑍subscript𝐻𝑅\displaystyle\sum_{\begin{subarray}{c}m\in A(B)\\ \alpha\beta\,\sigma\sigma^{\prime}\end{subarray}}{s}^{z}_{\sigma\sigma^{\prime% }}{\lambda^{A(B)}}_{\alpha\beta}\,{\hat{c}}^{\dagger}_{m\alpha\sigma}{\hat{c}}% _{m\beta\sigma^{\prime}}+H_{Z}+H_{R}∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_m ∈ italic_A ( italic_B ) end_CELL end_ROW start_ROW start_CELL italic_α italic_β italic_σ italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_σ italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT italic_A ( italic_B ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_α italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_m italic_β italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT (7)
+\displaystyle++ m,nA(B)αβσtαβA(B)mnc^mασc^nβσsubscript𝑚𝑛𝐴𝐵𝛼𝛽𝜎superscriptsubscript𝑡𝛼𝛽𝐴𝐵delimited-⟨⟩𝑚𝑛subscriptsuperscript^𝑐𝑚𝛼𝜎subscript^𝑐𝑛𝛽𝜎\displaystyle\sum_{\begin{subarray}{c}\langle m,n\rangle\in A(B)\\ \alpha\beta\,\sigma\end{subarray}}{t_{\alpha\beta}^{A(B)\langle mn\rangle}\,{% \hat{c}}^{\dagger}_{m\alpha\sigma}{\hat{c}}_{n\beta\sigma}}∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ⟨ italic_m , italic_n ⟩ ∈ italic_A ( italic_B ) end_CELL end_ROW start_ROW start_CELL italic_α italic_β italic_σ end_CELL end_ROW end_ARG end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A ( italic_B ) ⟨ italic_m italic_n ⟩ end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_α italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_n italic_β italic_σ end_POSTSUBSCRIPT (10)
+\displaystyle++ mA,nBαβσγ(tαβAmn+tαβBmn)c^mασc^nβσ+H.c.formulae-sequencesubscriptdelimited-⟨⟩formulae-sequence𝑚𝐴𝑛𝐵𝛼𝛽𝜎𝛾superscriptsubscript𝑡𝛼𝛽𝐴delimited-⟨⟩𝑚𝑛superscriptsubscript𝑡𝛼𝛽𝐵delimited-⟨⟩𝑚𝑛subscriptsuperscript^𝑐𝑚𝛼𝜎subscript^𝑐𝑛𝛽𝜎𝐻𝑐\displaystyle\sum_{\begin{subarray}{c}\langle m\in A,n\in B\rangle\\ \alpha\beta\,\sigma\end{subarray}}{\gamma\left(t_{\alpha\beta}^{A\langle mn% \rangle}+t_{\alpha\beta}^{B\langle mn\rangle}\right){\hat{c}}^{\dagger}_{m% \alpha\sigma}{\hat{c}}_{n\beta\sigma}}+H.c.∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ⟨ italic_m ∈ italic_A , italic_n ∈ italic_B ⟩ end_CELL end_ROW start_ROW start_CELL italic_α italic_β italic_σ end_CELL end_ROW end_ARG end_POSTSUBSCRIPT italic_γ ( italic_t start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A ⟨ italic_m italic_n ⟩ end_POSTSUPERSCRIPT + italic_t start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B ⟨ italic_m italic_n ⟩ end_POSTSUPERSCRIPT ) over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_α italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_n italic_β italic_σ end_POSTSUBSCRIPT + italic_H . italic_c . (13)

Indices {m,n}𝑚𝑛\{m,n\}{ italic_m , italic_n }, {σ,σ}𝜎superscript𝜎\{\sigma,\sigma^{\prime}\}{ italic_σ , italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT }, and {α,β}𝛼𝛽\{\alpha,\beta\}{ italic_α , italic_β } enumerate lattice sites, spins, and orbitals; e.g., operator c^mασsubscriptsuperscript^𝑐𝑚𝛼𝜎{\hat{c}}^{\dagger}_{m\alpha\sigma}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_α italic_σ end_POSTSUBSCRIPT (c^mασsubscript^𝑐𝑚𝛼𝜎{\hat{c}}_{m\alpha\sigma}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_m italic_α italic_σ end_POSTSUBSCRIPT) creates (annihilates) an electron with orbital α𝛼\alphaitalic_α and spin σ𝜎\sigmaitalic_σ at m𝑚mitalic_m-th lattice site. Since both materials have similar lattice constants, we assume the same value a=0.319𝑎0.319a=0.319italic_a = 0.319 nm on both sides. The lattice combined of two materials MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT (A) and WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT (B) is presented in Fig. 4(a). The potential energy of the electrostatic confinement at the m𝑚mitalic_m-th lattice site φm=|e|ϕ(xm,ym)subscript𝜑𝑚𝑒italic-ϕsubscript𝑥𝑚subscript𝑦𝑚\varphi_{m}=-|e|\phi(x_{m},y_{m})italic_φ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = - | italic_e | italic_ϕ ( italic_x start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) together with the on-site energies ϵαA(B)superscriptsubscriptitalic-ϵ𝛼𝐴𝐵\epsilon_{\alpha}^{A(B)}italic_ϵ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A ( italic_B ) end_POSTSUPERSCRIPT enter the diagonal matrix elements. λαβA(B)superscriptsubscript𝜆𝛼𝛽𝐴𝐵\lambda_{\alpha\beta}^{A(B)}italic_λ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A ( italic_B ) end_POSTSUPERSCRIPT express the intrinsic spin-orbit coupling [37] and szsuperscript𝑠𝑧s^{z}italic_s start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT stands for the z𝑧zitalic_z-Pauli-matrix. Both of them take different values at A𝐴Aitalic_A(B𝐵Bitalic_B) material sides. HZsubscript𝐻𝑍H_{Z}italic_H start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT stands for the Zeeman Hamiltonian, and HRsubscript𝐻𝑅H_{R}italic_H start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT for the Rashba spin-orbit term [11]. The hopping elements tαβA(B)mnsuperscriptsubscript𝑡𝛼𝛽𝐴𝐵delimited-⟨⟩𝑚𝑛t_{\alpha\beta}^{A(B)\langle mn\rangle}italic_t start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A ( italic_B ) ⟨ italic_m italic_n ⟩ end_POSTSUPERSCRIPT depend on the nearest neighbours mndelimited-⟨⟩𝑚𝑛\langle mn\rangle⟨ italic_m italic_n ⟩ link direction and if the hopping is inside the given material we assume its hopping parameter A𝐴Aitalic_A or B𝐵Bitalic_B (second to last element of Eq. 4). The situation is quite different for hopping between materials; in such cases we take the simple average of hoppings when crossing the junction, i.e., γ=12𝛾12\gamma=\frac{1}{2}italic_γ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [38] in the last element of Eq. 4.

We utilize a TB model which uses only three d𝑑ditalic_d metal-orbitals: dz2subscript𝑑superscript𝑧2d_{z^{2}}italic_d start_POSTSUBSCRIPT italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT, dxysubscript𝑑𝑥𝑦d_{xy}italic_d start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT, dx2y2subscript𝑑superscript𝑥2superscript𝑦2d_{x^{2}-y^{2}}italic_d start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT on a triangular lattice of Mo (or W atoms), with the nearest-neighbors hoppings [39]. This simple model can correctly represent the dispersion relation and the orbital composition close to the K𝐾Kitalic_K point in the Brillouin zone (BZ) near the band edges, where the Bloch states mainly consist of metal d𝑑ditalic_d orbitals [40]. Because in our calculations we are concerned solely with states derived from the minimum of the CB, at K,K𝐾superscript𝐾K,K^{\prime}italic_K , italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT points, the tight-binding model used is sufficient.

We assume an armchair interface between materials – see Fig. 4(a). Both termination types, armchair and zigzag are found to be stable in TMDC lateral heterostructures [16], however, zigzag termination, also analyzed by us, does not allow for proper valley-qubit transitions. This phenomenon will be analyzed in Section IV.

Refer to caption
Figure 4: (a) Studied nanoflake model with lateral heterostructure (located at xj=7subscript𝑥𝑗7x_{j}=7italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 7 nm) of two materials. (b) quantum dot eigenstates (colors denote localization of states in QD – yellow states strongly localize, while dark violet are in-gap edge states) with the lowest CB states (inset) forming spin-valley subspace from which we will select (upon applying B field) two states (|K|K\!\downarrow\rangle| italic_K ↓ ⟩, |K|K^{\prime}\!\downarrow\rangle| italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↓ ⟩) that form valley qubit. (c) Confinement potential profile within the fake lattice at initial step and after some confinement modulation that push the electron packet toward the junction barrier – visible in (d). (e) Perpendicular electric field that induce Rashba SOC. (f) Resonant confinement potential modulation induces intervalley transition gradually transforming density from K𝐾Kitalic_K to Ksuperscript𝐾K^{\prime}italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT valley.

II.3 Poisson-Schrödinger method

Voltages applied to the gates (relative to the substrate) are used to create the QD confinement potential in the monolayers area. Electric dipole charge at the junction is responsible for the built-in electric field. The confined electron itself also carries charge density (which we account for via the mean field approach). To calculate the realistic electrostatic potential ϕ(𝐫)italic-ϕ𝐫\phi(\mathbf{r})italic_ϕ ( bold_r ) we solve the generalized Poisson equation [41]:

(ε0ε(𝐫)Φ(𝐫))bold-∇subscript𝜀0𝜀𝐫bold-∇Φ𝐫\displaystyle\boldsymbol{\nabla}\!\cdot\left(\varepsilon_{0}\varepsilon(% \mathbf{r})\boldsymbol{\nabla}\Phi(\mathbf{r})\right)bold_∇ ⋅ ( italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ε ( bold_r ) bold_∇ roman_Φ ( bold_r ) ) =(ρe(𝐫)+d+(𝐫)+d(𝐫)),absentsubscript𝜌𝑒𝐫superscript𝑑𝐫superscript𝑑𝐫\displaystyle=-\left(\rho_{e}(\mathbf{r})+d^{+}\!(\mathbf{r})+d^{-}\!(\mathbf{% r})\right),= - ( italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) + italic_d start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( bold_r ) + italic_d start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( bold_r ) ) ,
ϕ(𝐫)italic-ϕ𝐫\displaystyle\phi(\mathbf{r})italic_ϕ ( bold_r ) =Φ(𝐫)ϕe(𝐫),absentΦ𝐫subscriptitalic-ϕ𝑒𝐫\displaystyle=\Phi(\mathbf{r})-\phi_{e}(\mathbf{r}),= roman_Φ ( bold_r ) - italic_ϕ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) , (14)

taking into account all of these components: voltages V{L,R,U,D}subscript𝑉LRUDV_{\mathrm{\{L,R,U,D\}}}italic_V start_POSTSUBSCRIPT { roman_L , roman_R , roman_U , roman_D } end_POSTSUBSCRIPT applied to the control gates G{L,R,U,D}subscriptGLRUD\mathrm{G}_{\mathrm{\{L,R,U,D\}}}roman_G start_POSTSUBSCRIPT { roman_L , roman_R , roman_U , roman_D } end_POSTSUBSCRIPT and to the highly doped substrate (kept at the referential potential V0=0subscript𝑉00V_{0}=0italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0), space-dependent permittivity ε(𝐫)𝜀𝐫\varepsilon(\mathbf{r})italic_ε ( bold_r ) of different materials in the device (we assume εMoS2=6.2subscript𝜀subscriptMoS26.2\varepsilon_{\mathrm{MoS_{2}}}=6.2italic_ε start_POSTSUBSCRIPT roman_MoS start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 6.2, εWS2=6.1subscript𝜀subscriptWS26.1\varepsilon_{\mathrm{WS_{2}}}=6.1italic_ε start_POSTSUBSCRIPT roman_WS start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 6.1, and εhBN=3.3subscript𝜀hBN3.3\varepsilon_{\mathrm{hBN}}=3.3italic_ε start_POSTSUBSCRIPT roman_hBN end_POSTSUBSCRIPT = 3.3 [42]), electric dipole charge d±(𝐫)superscript𝑑plus-or-minus𝐫d^{\pm}(\mathbf{r})italic_d start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT ( bold_r ), and electron charge density ρe(𝐫)subscript𝜌𝑒𝐫\rho_{e}(\mathbf{r})italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) itself. At the lateral and top sides of the computational box we apply Neumann boundary conditions with zeroing normal component of the electric field. To remove electron self-interaction we subtract electron potential itself ϕe(𝐫)subscriptitalic-ϕ𝑒𝐫\phi_{e}(\mathbf{r})italic_ϕ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) from the total potential Φ(𝐫)Φ𝐫\Phi(\mathbf{r})roman_Φ ( bold_r ). Electron potential ϕe(𝐫)subscriptitalic-ϕ𝑒𝐫\phi_{e}(\mathbf{r})italic_ϕ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) is calculated using the standard Poisson equation: 2ϕe(𝐫)=ρe(𝐫)/(ε0ε(𝐫))superscriptbold-∇2subscriptitalic-ϕ𝑒𝐫subscript𝜌𝑒𝐫subscript𝜀0𝜀𝐫\boldsymbol{\nabla}^{2}\phi_{e}(\mathbf{r})=-\rho_{e}(\mathbf{r})/(\varepsilon% _{0}\varepsilon(\mathbf{r}))bold_∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) = - italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) / ( italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ε ( bold_r ) ) without any voltage conditions on the gates or the substrate. Further details of the used method can be found in Ref. [10].

To calculate the electron eigenstates we solve the Schrödinger equation for the Hamiltonian (Eq. 4) with electrostatic potential ϕ(𝐫)italic-ϕ𝐫\phi(\mathbf{r})italic_ϕ ( bold_r ) which is calculated via the Poisson equation,

H[ϕ(𝐫)]ψn(𝐫)=Enψn(𝐫).𝐻delimited-[]italic-ϕ𝐫subscript𝜓𝑛𝐫subscript𝐸𝑛subscript𝜓𝑛𝐫H[\phi(\mathbf{r})]\psi_{n}(\mathbf{r})=E_{n}\psi_{n}(\mathbf{r}).italic_H [ italic_ϕ ( bold_r ) ] italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_r ) = italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_r ) . (15)

On the other hand, to calculate the electrostatic potential one must solve the Poisson equation, which in turn requires knowledge of the electron charge density in a given QD state ρe(𝐫)=|e||ψn(𝐫)|2subscript𝜌𝑒𝐫𝑒superscriptsubscript𝜓𝑛𝐫2\rho_{e}(\mathbf{r})=-|e||\psi_{n}(\mathbf{r})|^{2}italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( bold_r ) = - | italic_e | | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_r ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This means that both of the equations need to be solved self-consistently.

Presented in Fig. 3 electron densities (dashed curves) for various QD-confinement configurations shows ground state density |e||ψ0(𝐫)|2𝑒superscriptsubscript𝜓0𝐫2-|e||\psi_{0}(\mathbf{r})|^{2}- | italic_e | | italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT calculated via the Poisson-Schrödinger system defined in Eqs. II.3 and 15.

During the time-dependent calculations we will be working in the calculated eigenstate basis, in which the full time-dependent wave function is represented as a linear combination of N𝑁Nitalic_N previously calculated basis states ψnsubscript𝜓𝑛\psi_{n}italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT:

Ψ(𝐫,t)=ncn(t)ψn(𝐫)eıEnt,Ψ𝐫𝑡subscript𝑛subscript𝑐𝑛𝑡subscript𝜓𝑛𝐫superscript𝑒italic-ıPlanck-constant-over-2-pisubscript𝐸𝑛𝑡\Psi(\mathbf{r},t)=\sum_{n}c_{n}(t)\,\psi_{n}(\mathbf{r})e^{-\frac{\imath}{% \hbar}E_{n}t},roman_Ψ ( bold_r , italic_t ) = ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_r ) italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_ı end_ARG start_ARG roman_ℏ end_ARG italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT , (16)

together with time-dependent amplitudes cn(t)subscript𝑐𝑛𝑡c_{n}(t)italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) and corresponding eigenvalues Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. The time evolution is governed by the time-dependent Schrödinger equation:

ıtΨ(𝐫,t)=H(𝐫,t)Ψ(𝐫,t),italic-ıPlanck-constant-over-2-pi𝑡Ψ𝐫𝑡𝐻𝐫𝑡Ψ𝐫𝑡\imath\hbar\frac{\partial}{\partial t}\Psi(\mathbf{r},t)=H(\mathbf{r},t)\Psi(% \mathbf{r},t),italic_ı roman_ℏ divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG roman_Ψ ( bold_r , italic_t ) = italic_H ( bold_r , italic_t ) roman_Ψ ( bold_r , italic_t ) , (17)

with the time-dependent Hamiltonian being a sum of the stationary part (Eq. 4) and the variable contribution to the potential energy:

H(𝐫,t)=H(𝐫)|e|δϕ(𝐫,t).𝐻𝐫𝑡𝐻𝐫𝑒𝛿italic-ϕ𝐫𝑡H(\mathbf{r},t)=H(\mathbf{r})-|e|\delta\phi(\mathbf{r},t).italic_H ( bold_r , italic_t ) = italic_H ( bold_r ) - | italic_e | italic_δ italic_ϕ ( bold_r , italic_t ) . (18)

The full time-dependent potential ϕ(𝐫,t)=ϕ(𝐫)+δϕ(𝐫,t)italic-ϕ𝐫𝑡italic-ϕ𝐫𝛿italic-ϕ𝐫𝑡\phi(\mathbf{r},t)=\phi(\mathbf{r})+\delta\phi(\mathbf{r},t)italic_ϕ ( bold_r , italic_t ) = italic_ϕ ( bold_r ) + italic_δ italic_ϕ ( bold_r , italic_t ) contains variable part δϕ(𝐫,t)𝛿italic-ϕ𝐫𝑡\delta\phi(\mathbf{r},t)italic_δ italic_ϕ ( bold_r , italic_t ), generated by modulation of the gate voltages. It is calculated by solving the Poisson equation for the variable density ρ(𝐫,t)𝜌𝐫𝑡\rho(\mathbf{r},t)italic_ρ ( bold_r , italic_t ) at every time step. Note that the charge density originates from the actual wave-function, thus the Schrödinger and Poisson equations are solved in a self-consistent way for each time step.

Insertion of (16) to the Schrödinger equation (17) gives a system of equations for time-derivatives of the expansion coefficients at subsequent moments in time:

c˙m(t)=ıncn(t)δmn(t)eı(EmEn)t.subscript˙𝑐𝑚𝑡italic-ıPlanck-constant-over-2-pisubscript𝑛subscript𝑐𝑛𝑡subscript𝛿𝑚𝑛𝑡superscript𝑒italic-ıPlanck-constant-over-2-pisubscript𝐸𝑚subscript𝐸𝑛𝑡\dot{c}_{m}(t)=-\frac{\imath}{\hbar}\sum_{n}c_{n}(t)\,\delta_{mn}(t)\,e^{\frac% {\imath}{\hbar}(E_{m}-E_{n})t}.over˙ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_t ) = - divide start_ARG italic_ı end_ARG start_ARG roman_ℏ end_ARG ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) italic_δ start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT ( italic_t ) italic_e start_POSTSUPERSCRIPT divide start_ARG italic_ı end_ARG start_ARG roman_ℏ end_ARG ( italic_E start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT . (19)

The matrix elements δmn(t)=|e|ψm|δϕ(𝐫,t)|ψnsubscript𝛿𝑚𝑛𝑡𝑒quantum-operator-productsubscript𝜓𝑚𝛿italic-ϕ𝐫𝑡subscript𝜓𝑛\delta_{mn}(t)=-|e|\langle\psi_{m}|\delta\phi(\mathbf{r},t)|\psi_{n}\rangleitalic_δ start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT ( italic_t ) = - | italic_e | ⟨ italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT | italic_δ italic_ϕ ( bold_r , italic_t ) | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ need to be calculated at every time step due to changes in the potential.

III Valley qubit manipulation

After describing the computational method, let us apply voltages to the device gates VU=VD=2.5subscript𝑉Usubscript𝑉D2.5V_{\mathrm{U}}=V_{\mathrm{D}}=-2.5italic_V start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT = - 2.5 V, VL=1.2subscript𝑉L1.2V_{\mathrm{L}}=-1.2italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT = - 1.2 V, VR=2subscript𝑉R2V_{\mathrm{R}}=-2italic_V start_POSTSUBSCRIPT roman_R end_POSTSUBSCRIPT = - 2 V and calculate self-consistently (via the Poisson-Schrödinger method) the QD confinement and electron ground state presented in Fig. 4(c,d)-left. Next, the time-dependent calculation starts. We change the left gate voltage to VL=1.5subscript𝑉L1.5V_{\mathrm{L}}=-1.5italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT = - 1.5 V, moving the QD confining potential (and electron density within) slightly to the flake center – Fig. 4(c,d)-center. Then we start the pumping process by applying oscillatory voltages to the left gate: VL(t)=1.50.5(1cos(ωt)/2V_{\mathrm{L}}(t)=-1.5-0.5(1\!-\!\cos(\omega t)/2italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT ( italic_t ) = - 1.5 - 0.5 ( 1 - roman_cos ( italic_ω italic_t ) / 2 V. Oscillatory VL(t)subscript𝑉L𝑡V_{\mathrm{L}}(t)italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT ( italic_t ) moves the electron back and forth towards the junction interface. Respective potential profiles along the x𝑥xitalic_x-axis for different VLsubscript𝑉LV_{\mathrm{L}}italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT voltages are also presented in Fig. 3(a).

Refer to caption
Figure 5: Dipole moment-induced intervalley transitions at the heterojunction interface for different frequencies ω𝜔\omegaitalic_ω of pumping that move the QD electron back and forth towards the junction. Transitions have a resonant nature, i.e., only a resonance period of T=4.5𝑇4.5T=4.5italic_T = 4.5 ps gives full transitions.

When the pumping frequency ω𝜔\omegaitalic_ω is tuned to the qubit states’ splitting (here, 2π/ω=4.52𝜋𝜔4.52\pi/\omega=4.52 italic_π / italic_ω = 4.5 ps), we start to observe the intervalley transitions presented in Fig. 5. At t=140𝑡140t=140italic_t = 140 ps, the valley isospin flips from 𝒦=1𝒦1\mathcal{K}=-1caligraphic_K = - 1 (K𝐾Kitalic_K valley) to 𝒦=1𝒦1\mathcal{K}=1caligraphic_K = 1 (Ksuperscript𝐾K^{\prime}italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT valley). These transitions are resonant and have the character of Rabi oscillations with incomplete transitions for out-of-resonance pumping frequency. The valley isospin value can be obtained by calculating the Fourier transform of the actual Ψ(𝐫,t)Ψ𝐫𝑡\Psi(\mathbf{r},t)roman_Ψ ( bold_r , italic_t ) wavefunction, as presented in Fig. 4(f) – details can be found in [10]. Resonant pumping gradually transfers density from K𝐾Kitalic_K to Ksuperscript𝐾K^{\prime}italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT valley within the hexagonal BZ (white dashed line).

The frequency needed to address the qubit states’ energy splitting of the order of meV reaches hundreds of GHz which might be problematic in any experimental setup. Luckily MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT has relatively low spin-orbit CB energy splitting which combined with properly tuned perpendicular magnetic field may lead to much smaller splitting in given spin subspace (e.g. {|K,|K}\{|K\downarrow\rangle,|K^{\prime}\downarrow\rangle\}{ | italic_K ↓ ⟩ , | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↓ ⟩ }). This in turn can be addressed by much smaller pumping frequencies. Such situation occurs next to levels crossing about B=2.4𝐵2.4B=2.4italic_B = 2.4 T shown in Fig. 10(b). The low CB spin-orbit splitting in MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT among all TMDC materials is the reason why it seems to be the most suitable for the proposed setup.

Refer to caption
Figure 6: Intervalley transitions in the regime of weaker electron density overlapping with the junction interface, tuned by adjusting the VL,minsubscript𝑉LminV_{\mathrm{L,min}}italic_V start_POSTSUBSCRIPT roman_L , roman_min end_POSTSUBSCRIPT value. Inset: resonance curves for different pumping frequency/period.

To get transitions, the dipole moment alone is insufficient; the heterojunction interface must have proper termination. For a junction oriented with a zigzag interface we have not obtained transitions. All the results presented in Figs. 5 and 6 were obtained for armchair termination between the MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT and WS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT monolayers. This problem will be analyzed in Section IV.

When the VL(t)subscript𝑉L𝑡V_{\mathrm{L}}(t)italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT ( italic_t ) minimum value is higher (smaller VL(t)subscript𝑉L𝑡V_{\mathrm{L}}(t)italic_V start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT ( italic_t ) oscillation amplitude) the junction’s ability to induce transitions between valleys decreases. Presented in Fig 6 are intervalley transitions with higher values of VL,min>2subscript𝑉Lmin2V_{\mathrm{L,min}}>-2italic_V start_POSTSUBSCRIPT roman_L , roman_min end_POSTSUBSCRIPT > - 2 V, resulting in weaker overlapping with the junction interface and a longer transition time (Rabi period) >140absent140>140> 140 ps. The resonant frequency is also slightly smaller (higher resonance period), and thus the resonance peak width is slightly reduced – see inset in Fig 6.

One should also note that in principle another definition of the qubit is possible [14], i.e. using the two lowest spin-valley states. However, such configuration needs to have some spin-flip mechanism also present in the system, e.g. in a form of the Rashba coupling which is also included in our numerical model, in the Hamiltonian (Eq. 4).

IV Intervalley transition conditions

There are two crucial conditions needed for intervalley transitions. The first one is related to the length scale of the linear dipole density at the junction interface, and the second one to the linear dipole orientation with respect to the monolayer lattice.

To meet the first condition, the perturbation, here in the form of the dipole potential, length scale should be comparable with the KK𝐾superscript𝐾K-K^{\prime}italic_K - italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT difference in the momentum space [14, 29].

Refer to caption
Figure 7: Fourier transform of the confinement potential with and without the junction potential (linear dipole).

This condition can be easily verified by calculating a Fourier transformation F𝐹Fitalic_F of the confinement φ(x,y)=|e|ϕ(x,y)𝜑𝑥𝑦𝑒italic-ϕ𝑥𝑦\varphi(x,y)=-|e|\phi(x,y)italic_φ ( italic_x , italic_y ) = - | italic_e | italic_ϕ ( italic_x , italic_y ). Presented in Fig. 7 is the squared modulus of the 2D Fourier transform |F[φ](kx,ky)|2superscript𝐹delimited-[]𝜑subscript𝑘𝑥subscript𝑘𝑦2|F[\varphi](k_{x},k_{y})|^{2}| italic_F [ italic_φ ] ( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT along the ky=0subscript𝑘𝑦0k_{y}=0italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 0 line (KK=(4π3a,0)superscript𝐾𝐾4𝜋3𝑎0K^{\prime}-K=(\frac{4\pi}{3a},0)italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_K = ( divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG , 0 ) – see the BZ in Fig. 4(f), and the lattice constant a=0.319𝑎0.319a=0.319italic_a = 0.319 nm) for the original confinement potential (blue curve), and the same potential but with the junction (dipole) potential subtracted (orange). The latter one simply contains the gate-defined QD potential alone. It is clearly visible that amplitude at kx=4π3a13.13subscript𝑘𝑥4𝜋3𝑎similar-to-or-equals13.13k_{x}=\frac{4\pi}{3a}\simeq 13.13italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG ≃ 13.13 nm11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT is almost an order of magnitude larger for the case with a junction potential (more high-k𝑘kitalic_k-vector components due to rapidly varying area of the linear dipole) than in the case without a junction (only slowly varying QD potential).

The second condition states that the dipole orientation with respect to a monolayer lattice should be of armchair type.

Refer to caption
Figure 8: (top) Various locations of the model Gaussian QD potential determine different positions of qubit state confined within. (bottom) Various directions of the Gaussian-shaped line potential that simulates the junction potential.

To study this problem we calculated potential matrix elements between the two basis states of the valley qubit: K|φ|Kquantum-operator-product𝐾𝜑superscript𝐾\langle{}K|\varphi|K^{\prime}\rangle⟨ italic_K | italic_φ | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩. To examine different junction orientations and qubit state positions with respect to the junction location, we assumed a model Gaussian-like potential and placed it in different positions, then we calculated eigenstates (using Eq. 4, but assuming MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT material for the whole flake) and take two states |Kket𝐾|K\rangle| italic_K ⟩ and |Kketsuperscript𝐾|K^{\prime}\rangle| italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ from the CB minimum (with the same spin, e.g. oriented upwards) for three different locations (A,B,C) – their densities are presented in Fig. 8(top). Next, we define a Gaussian-shaped linear potential φ𝜑\varphiitalic_φ that simulates the junction oriented along different directions θ𝜃\thetaitalic_θ – as presented in Fig. 8(bottom). The potential is positioned along a line passing through the flake center, but it can also be shifted laterally by some displacement δ𝛿\deltaitalic_δ.

Refer to caption
Figure 9: Matrix elements of artificially designed potential that mimics junction potential (oriented along different directions θ𝜃\thetaitalic_θ and with possible lateral shift δ𝛿\deltaitalic_δ) for valley qubit states |Kket𝐾|K\rangle| italic_K ⟩ and |Kketsuperscript𝐾|K^{\prime}\rangle| italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ located at three different positions (A,B,C).

Finally, we calculated the matrix elements K|φ|Kquantum-operator-product𝐾𝜑superscript𝐾\langle{}K|\varphi|K^{\prime}\rangle⟨ italic_K | italic_φ | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ for different junction orientations and qubit state positions.

Presented in Fig. 9(top) is the modulus of the matrix elements as a function of artificial junction orientation θ𝜃\thetaitalic_θ for various state positions. It is clearly visible that nonzero elements appear only next to θ=π6,π2,5π6𝜃𝜋6𝜋25𝜋6\theta=\frac{\pi}{6},\frac{\pi}{2},\frac{5\pi}{6}italic_θ = divide start_ARG italic_π end_ARG start_ARG 6 end_ARG , divide start_ARG italic_π end_ARG start_ARG 2 end_ARG , divide start_ARG 5 italic_π end_ARG start_ARG 6 end_ARG, i.e. for a junction interface along the armchair termination. For example for θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG we have the same termination as in Fig. 4(a). On the other hand, for zigzag terminations, i.e. θ=0,π3,2π3𝜃0𝜋32𝜋3\theta=0,\frac{\pi}{3},\frac{2\pi}{3}italic_θ = 0 , divide start_ARG italic_π end_ARG start_ARG 3 end_ARG , divide start_ARG 2 italic_π end_ARG start_ARG 3 end_ARG, the matrix elements are zero for any state positions A-C. Also note that for θ=5π6𝜃5𝜋6\theta=\frac{5\pi}{6}italic_θ = divide start_ARG 5 italic_π end_ARG start_ARG 6 end_ARG, i.e. when states A-C are oriented along the junction, matrix elements are the same regardless of state’s position. Our findings that armchair termination may mix different valleys are in agreement with the present literature. Armchair edges mix valleys and induce transitions in graphene-like structures [43, 44, 45, 46]. Same edge-dependent valley mixing was proven for MoS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT nanoribbons [47], or more generally for TMDC monolayers [28].

Armchair terminations give nonzero matrix elements for zero (δ=0𝛿0\delta=0italic_δ = 0), or small nonzero (δ=a/2𝛿𝑎2\delta=a/2italic_δ = italic_a / 2) shifts as presented in Figs. 9(middle,bottom). It is also characteristic that detailed values of the matrix elements (real and imaginary part) strongly depend on the shift parameter, and for nonzero δ=a/2𝛿𝑎2\delta=a/2italic_δ = italic_a / 2 they have different form for different angles θ=π6𝜃𝜋6\theta=\frac{\pi}{6}italic_θ = divide start_ARG italic_π end_ARG start_ARG 6 end_ARG and θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG – especially imaginary parts are opposite – see Fig. 9(bottom). For δ=0𝛿0\delta=0italic_δ = 0 this dependence is a bit weaker but still visible, especially for the B and C state positions – Fig. 9(middle).

IV.1 Analytical model for intervalley coupling

To get better understanding of the underlying physics we also built a simple analytical model that captures only the features important to qualitatively describe the intervalley transitions mechanism – sharp elongated potential aligned along a given direction with respect to the lattice vectors. Then we developed analytical formula for K|φ|Kquantum-operator-product𝐾𝜑superscript𝐾\langle{}K|\varphi|K^{\prime}\rangle⟨ italic_K | italic_φ | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ element in that case. However, in order to estimate the detailed operation of the device, such as the optimal gate layout or the voltages needed to effectively control the qubit or its timings, it is necessary to get back to the original Poisson-tight-binding model.

To build the analytic model we assume 2D-Gaussian-like potential, centered at 𝐱0=(x0,y0)Tsubscript𝐱0superscriptsubscript𝑥0subscript𝑦0𝑇\mathbf{x}_{0}=(x_{0},y_{0})^{T}bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT and rotated by θ𝜃\thetaitalic_θ, in a form:

φG(𝐱,𝐱0,θ,sx,sy)=U0e12(𝐱𝐱0)T𝐀(𝐱𝐱0),subscript𝜑𝐺𝐱subscript𝐱0𝜃subscript𝑠𝑥subscript𝑠𝑦subscript𝑈0superscript𝑒12superscript𝐱subscript𝐱0𝑇𝐀𝐱subscript𝐱0\displaystyle\varphi_{G}(\mathbf{x},\mathbf{x}_{0},\theta,s_{x},s_{y})=U_{0}e^% {-\frac{1}{2}(\mathbf{x}-\mathbf{x}_{0})^{T}\!\mathbf{A}(\mathbf{x}-\mathbf{x}% _{0})},italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( bold_x , bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ , italic_s start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_s start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) = italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( bold_x - bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_A ( bold_x - bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT ,
𝐀=(cos2θsx2+sin2θsy2sin2θsx2/2+sin2θsy2/2sin2θsx2/2+sin2θsy2/2sin2θsx2+cos2θsy2)𝐀matrixsuperscript2𝜃subscriptsuperscript𝑠2𝑥superscript2𝜃subscriptsuperscript𝑠2𝑦2𝜃subscriptsuperscript𝑠2𝑥22𝜃subscriptsuperscript𝑠2𝑦22𝜃subscriptsuperscript𝑠2𝑥22𝜃subscriptsuperscript𝑠2𝑦2superscript2𝜃subscriptsuperscript𝑠2𝑥superscript2𝜃subscriptsuperscript𝑠2𝑦\displaystyle\mathbf{A}=\begin{pmatrix}\frac{\cos^{2}\theta}{s^{2}_{x}}+\frac{% \sin^{2}\theta}{s^{2}_{y}}&-\frac{\sin 2\theta}{s^{2}_{x}/2}+\frac{\sin 2% \theta}{s^{2}_{y}/2}\\ -\frac{\sin 2\theta}{s^{2}_{x}/2}+\frac{\sin 2\theta}{s^{2}_{y}/2}&\frac{\sin^% {2}\theta}{s^{2}_{x}}+\frac{\cos^{2}\theta}{s^{2}_{y}}\end{pmatrix}bold_A = ( start_ARG start_ROW start_CELL divide start_ARG roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_ARG + divide start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_ARG end_CELL start_CELL - divide start_ARG roman_sin 2 italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT / 2 end_ARG + divide start_ARG roman_sin 2 italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT / 2 end_ARG end_CELL end_ROW start_ROW start_CELL - divide start_ARG roman_sin 2 italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT / 2 end_ARG + divide start_ARG roman_sin 2 italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT / 2 end_ARG end_CELL start_CELL divide start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_ARG + divide start_ARG roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ end_ARG start_ARG italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_ARG end_CELL end_ROW end_ARG ) , (22)

with 𝐱=(x,y)𝐱𝑥𝑦\mathbf{x}=(x,y)bold_x = ( italic_x , italic_y ) and standard deviations in both orthogonal directions defined as (sz,sy)subscript𝑠𝑧subscript𝑠𝑦(s_{z},s_{y})( italic_s start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_s start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ).

Now to calculate the Fourier transform of the φGsubscript𝜑𝐺\varphi_{G}italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT potential on a triangular lattice of Mo (or W) atoms, defined by lattice vectors 𝐚1=a(1,0)subscript𝐚1𝑎10\mathbf{a}_{1}=a(1,0)bold_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_a ( 1 , 0 ) and 𝐚2=a2(1,3)subscript𝐚2𝑎213\mathbf{a}_{2}=\frac{a}{2}(1,\sqrt{3})bold_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG italic_a end_ARG start_ARG 2 end_ARG ( 1 , square-root start_ARG 3 end_ARG ), we have to perform the following summation over the whole lattice:

φG(𝐪,𝐤)=1NUCj1,j2ei(𝐤𝐪)(j1𝐚1+j2𝐚2)φG(j1𝐚1+j2𝐚2),subscript𝜑𝐺𝐪𝐤1subscript𝑁𝑈𝐶subscriptsubscript𝑗1subscript𝑗2superscript𝑒𝑖𝐤𝐪subscript𝑗1subscript𝐚1subscript𝑗2subscript𝐚2subscript𝜑𝐺subscript𝑗1subscript𝐚1subscript𝑗2subscript𝐚2\varphi_{G}(\mathbf{q},\mathbf{k})=\frac{1}{N_{UC}}\sum_{j_{1},j_{2}}e^{i(% \mathbf{k}-\mathbf{q})(j_{1}\mathbf{a}_{1}+j_{2}\mathbf{a}_{2})}\varphi_{G}(j_% {1}\mathbf{a}_{1}+j_{2}\mathbf{a}_{2}),italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( bold_q , bold_k ) = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_U italic_C end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_j start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i ( bold_k - bold_q ) ( italic_j start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT bold_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT bold_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_j start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT bold_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT bold_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) , (23)

with NUCsubscript𝑁𝑈𝐶N_{UC}italic_N start_POSTSUBSCRIPT italic_U italic_C end_POSTSUBSCRIPT unit cells. For a finite lattice this sum is hard to be evaluated analytically, however, if one extend the system to infinite lattice the sum (23) can be approximated in continuum limit as (𝐐=𝐤𝐪𝐐𝐤𝐪\mathbf{Q}=\mathbf{k}-\mathbf{q}bold_Q = bold_k - bold_q):

φG(𝐐)𝑑j1𝑑j2eia(Qx(j1+j2/2)+Qyj23/2)similar-to-or-equalssubscript𝜑𝐺𝐐differential-dsubscript𝑗1differential-dsubscript𝑗2superscript𝑒𝑖𝑎subscript𝑄𝑥subscript𝑗1subscript𝑗22subscript𝑄𝑦subscript𝑗232\displaystyle\varphi_{G}(\mathbf{Q})\simeq\int dj_{1}dj_{2}e^{ia\left(Q_{x}(j_% {1}+j_{2}/2)+Q_{y}j_{2}\sqrt{3}/2\right)}italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( bold_Q ) ≃ ∫ italic_d italic_j start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_d italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_a ( italic_Q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_j start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / 2 ) + italic_Q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT square-root start_ARG 3 end_ARG / 2 ) end_POSTSUPERSCRIPT ×\displaystyle\times×
×φG(a(j1+j2/2),aj23/2),x0,y0)\displaystyle\times\varphi_{G}(a(j_{1}+j_{2}/2),aj_{2}\sqrt{3}/2),x_{0},y_{0})× italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_a ( italic_j start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / 2 ) , italic_a italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT square-root start_ARG 3 end_ARG / 2 ) , italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) . (24)

The integral (IV.1) can be calculated using formula for generalized Gaussian integral (here for 2D case, (j1,j2)subscript𝑗1subscript𝑗2(j_{1},j_{2})( italic_j start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_j start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT )). After tedious calculations, and assuming that the potential φGsubscript𝜑𝐺\varphi_{G}italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT is centered at the position (x0,y0)=(δsinθ,δcosθ)subscript𝑥0subscript𝑦0𝛿𝜃𝛿𝜃(x_{0},y_{0})=(-\delta\sin\theta,\delta\cos\theta)( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = ( - italic_δ roman_sin italic_θ , italic_δ roman_cos italic_θ ) where δ𝛿\deltaitalic_δ is the lateral displacement, one arrives at the formula:

φG(𝐐)similar-to-or-equalssubscript𝜑𝐺𝐐absent\displaystyle\varphi_{G}(\mathbf{Q})\simeqitalic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( bold_Q ) ≃
UGe14((sx2+sy2)(Qx2+Qy2)+(sx2sy2)((Qx2Qy2)cos2θ+2QxQysin2θ))subscript𝑈𝐺superscript𝑒14subscriptsuperscript𝑠2𝑥subscriptsuperscript𝑠2𝑦subscriptsuperscript𝑄2𝑥subscriptsuperscript𝑄2𝑦subscriptsuperscript𝑠2𝑥subscriptsuperscript𝑠2𝑦subscriptsuperscript𝑄2𝑥subscriptsuperscript𝑄2𝑦2𝜃2subscript𝑄𝑥subscript𝑄𝑦2𝜃\displaystyle U_{G}e^{-\frac{1}{4}\left((s^{2}_{x}+s^{2}_{y})(Q^{2}_{x}+Q^{2}_% {y})+(s^{2}_{x}-s^{2}_{y})\left((Q^{2}_{x}-Q^{2}_{y})\cos 2\theta+2Q_{x}Q_{y}% \sin 2\theta\right)\right)}italic_U start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG ( ( italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) + ( italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) ( ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) roman_cos 2 italic_θ + 2 italic_Q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT roman_sin 2 italic_θ ) ) end_POSTSUPERSCRIPT
×eiδ(QxsinθQycosθ),absentsuperscript𝑒𝑖𝛿subscript𝑄𝑥𝜃subscript𝑄𝑦𝜃\displaystyle\times e^{-i\delta\left(Q_{x}\sin\theta-Q_{y}\cos\theta\right)},× italic_e start_POSTSUPERSCRIPT - italic_i italic_δ ( italic_Q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_sin italic_θ - italic_Q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT roman_cos italic_θ ) end_POSTSUPERSCRIPT , (25)

with Ug=4π3a2sxsyU0subscript𝑈𝑔4𝜋3superscript𝑎2subscript𝑠𝑥subscript𝑠𝑦subscript𝑈0U_{g}=\frac{4\pi}{\sqrt{3}a^{2}}s_{x}s_{y}U_{0}italic_U start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = divide start_ARG 4 italic_π end_ARG start_ARG square-root start_ARG 3 end_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_s start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. If we lastly assume that the potential is significantly elongated in x𝑥xitalic_x direction, i.e. sxsymuch-greater-thansubscript𝑠𝑥subscript𝑠𝑦s_{x}\gg s_{y}italic_s start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ≫ italic_s start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT (before rotation by θ𝜃\thetaitalic_θ) we finally get:

φG(𝐐)UGe14(sx2(Qx2+Qy2)+sx2((Qx2Qy2)cos2θ+2QxQysin2θ))similar-to-or-equalssubscript𝜑𝐺𝐐subscript𝑈𝐺superscript𝑒14subscriptsuperscript𝑠2𝑥subscriptsuperscript𝑄2𝑥subscriptsuperscript𝑄2𝑦subscriptsuperscript𝑠2𝑥subscriptsuperscript𝑄2𝑥subscriptsuperscript𝑄2𝑦2𝜃2subscript𝑄𝑥subscript𝑄𝑦2𝜃\displaystyle\varphi_{G}(\mathbf{Q})\simeq U_{G}e^{-\frac{1}{4}\left(s^{2}_{x}% (Q^{2}_{x}+Q^{2}_{y})+s^{2}_{x}\left((Q^{2}_{x}-Q^{2}_{y})\cos 2\theta+2Q_{x}Q% _{y}\sin 2\theta\right)\right)}italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( bold_Q ) ≃ italic_U start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG ( italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) + italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) roman_cos 2 italic_θ + 2 italic_Q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT roman_sin 2 italic_θ ) ) end_POSTSUPERSCRIPT
×eiδ(QxsinθQycosθ).absentsuperscript𝑒𝑖𝛿subscript𝑄𝑥𝜃subscript𝑄𝑦𝜃\displaystyle\times e^{-i\delta\left(Q_{x}\sin\theta-Q_{y}\cos\theta\right)}.× italic_e start_POSTSUPERSCRIPT - italic_i italic_δ ( italic_Q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_sin italic_θ - italic_Q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT roman_cos italic_θ ) end_POSTSUPERSCRIPT . (26)

The characteristic feature of φG(𝐐)subscript𝜑𝐺𝐐\varphi_{G}(\mathbf{Q})italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( bold_Q ) is that it is elongated, in 𝐐𝐐\mathbf{Q}bold_Q space, in direction θ+π/2𝜃𝜋2\theta+\pi/2italic_θ + italic_π / 2, perpendicular to φG(𝐱)subscript𝜑𝐺𝐱\varphi_{G}(\mathbf{\mathbf{x}})italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( bold_x ) in real space. If we now calculate the matrix element K|φG|Kquantum-operator-product𝐾subscript𝜑𝐺superscript𝐾\langle K|\varphi_{G}|K^{\prime}\rangle⟨ italic_K | italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ by putting 𝐐=KK=(4π3a,0)𝐐superscript𝐾𝐾4𝜋3𝑎0\mathbf{Q}=K^{\prime}-K=(\frac{4\pi}{3a},0)bold_Q = italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_K = ( divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG , 0 ) we get a value which takes maximum at θ=π2𝜃𝜋2\theta=-\frac{\pi}{2}italic_θ = - divide start_ARG italic_π end_ARG start_ARG 2 end_ARG, i.e. the armchair direction:

K|φG(θ=π/2)|K=UGeiδ4π3a.quantum-operator-product𝐾subscript𝜑𝐺𝜃𝜋2superscript𝐾subscript𝑈𝐺superscript𝑒𝑖𝛿4𝜋3𝑎\langle K|\varphi_{G}(\theta=-\pi/2)|K^{\prime}\rangle=U_{G}e^{i\delta\frac{4% \pi}{3a}}.⟨ italic_K | italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_θ = - italic_π / 2 ) | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ = italic_U start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_δ divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG end_POSTSUPERSCRIPT . (27)

Same values can be obtained for two other armchair directions, i.e. θ=π6𝜃𝜋6\theta=\frac{\pi}{6}italic_θ = divide start_ARG italic_π end_ARG start_ARG 6 end_ARG and 5π65𝜋6\frac{5\pi}{6}divide start_ARG 5 italic_π end_ARG start_ARG 6 end_ARG:

K|φG(θ=π/6)|K=K|φG(θ=5π/6)|K=UGeiδ4π3a.quantum-operator-product𝐾subscript𝜑𝐺𝜃𝜋6superscript𝐾quantum-operator-product𝐾subscript𝜑𝐺𝜃5𝜋6superscript𝐾subscript𝑈𝐺superscript𝑒𝑖𝛿4𝜋3𝑎\langle K|\varphi_{G}(\theta=\pi/6)|K^{\prime}\rangle=\langle K|\varphi_{G}(% \theta=5\pi/6)|K^{\prime}\rangle=U_{G}e^{i\delta\frac{4\pi}{3a}}.⟨ italic_K | italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_θ = italic_π / 6 ) | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ = ⟨ italic_K | italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_θ = 5 italic_π / 6 ) | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ = italic_U start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_δ divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG end_POSTSUPERSCRIPT . (28)

These results are in agreement with numerical estimations from Fig. 9. If we then construct a junction to have a form similar to one from Fig. 10(a) we get:

K|φG(θ=π/6)+φG(θ=π/2)|Ksimilar-toquantum-operator-product𝐾subscript𝜑𝐺𝜃𝜋6subscript𝜑𝐺𝜃𝜋2superscript𝐾absent\displaystyle\langle K|\varphi_{G}(\theta=\pi/6)+\varphi_{G}(\theta=\pi/2)|K^{% \prime}\rangle\sim⟨ italic_K | italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_θ = italic_π / 6 ) + italic_φ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_θ = italic_π / 2 ) | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ∼
ψπ6eiδ4π3a+ψπ2eiδ4π3a,similar-toabsentsubscript𝜓𝜋6superscript𝑒𝑖𝛿4𝜋3𝑎subscript𝜓𝜋2superscript𝑒𝑖𝛿4𝜋3𝑎\displaystyle\sim\psi_{\frac{\pi}{6}}e^{i\delta\frac{4\pi}{3a}}+\psi_{\frac{% \pi}{2}}e^{-i\delta\frac{4\pi}{3a}},∼ italic_ψ start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 6 end_ARG end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_δ divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG end_POSTSUPERSCRIPT + italic_ψ start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_δ divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG end_POSTSUPERSCRIPT , (29)

where ψπ6subscript𝜓𝜋6\psi_{\frac{\pi}{6}}italic_ψ start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 6 end_ARG end_POSTSUBSCRIPT is the amplitude for finding electron next to one arm (for θ=π/6𝜃𝜋6\theta=\pi/6italic_θ = italic_π / 6) of the junction, and ψπ2subscript𝜓𝜋2\psi_{\frac{\pi}{2}}italic_ψ start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT for the other. Now assume that the displacement δ𝛿\deltaitalic_δ is smaller than the lattice vector a𝑎aitalic_a then the Hamiltonian in {K,K}𝐾superscript𝐾\{K,K^{\prime}\}{ italic_K , italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT } basis takes the form:

HKKψπ6(τxKδτy)+ψπ2(τx+Kδτy),similar-tosubscript𝐻𝐾superscript𝐾subscript𝜓𝜋6subscript𝜏𝑥𝐾𝛿subscript𝜏𝑦subscript𝜓𝜋2subscript𝜏𝑥𝐾𝛿subscript𝜏𝑦H_{KK^{\prime}}\sim\psi_{\frac{\pi}{6}}(\tau_{x}-K\delta\tau_{y})+\psi_{\frac{% \pi}{2}}(\tau_{x}+K\delta\tau_{y}),italic_H start_POSTSUBSCRIPT italic_K italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∼ italic_ψ start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 6 end_ARG end_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - italic_K italic_δ italic_τ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) + italic_ψ start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_K italic_δ italic_τ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) , (30)

with K=4π3a𝐾4𝜋3𝑎K=\frac{4\pi}{3a}italic_K = divide start_ARG 4 italic_π end_ARG start_ARG 3 italic_a end_ARG and τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT being valley Pauli matrices. What is obvious, operations connected to electron interaction with junction arms do not commute.

V Valley geometric phase

Results from Fig. 9 motivated us to investigate the possibility of valley qubit manipulation via a geometric phase. Berry phase [48] naturally emerges in non-degenerate quantum state upon cyclic, adiabatic manipulation. It can also be generalized on degenerate systems [49] leading to so-called non-Abelian geometric phase. Such non-commutativity of spin rotation matrices, connected to adiabatic manipulation of electron spin, when it moves along a closed path, leads to an effective spin manipulation scheme [50, 51, 52]. This scheme utilizes spin-orbit coupling, present in TMDC or III-V materials, and does not need any external magnetic field (degeneracy), which may limit qubit scalability. Resulting spin rotations also do not depend on dynamic evolution details, but only on the geometry of a given closed path.

When analysing results from Fig. 9, it became clear to us that intervalley matrix elements K|φ|Kquantum-operator-product𝐾𝜑superscript𝐾\langle{}K|\varphi|K^{\prime}\rangle⟨ italic_K | italic_φ | italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ for different junction angles (i.e., θ=π6𝜃𝜋6\theta=\frac{\pi}{6}italic_θ = divide start_ARG italic_π end_ARG start_ARG 6 end_ARG and θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG) may be interpreted as (compare middle panels and position C) Tπ6=η2(τxτy)subscript𝑇𝜋6𝜂2subscript𝜏𝑥subscript𝜏𝑦T_{\frac{\pi}{6}}=\frac{\eta}{2}(\tau_{x}-\tau_{y})italic_T start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 6 end_ARG end_POSTSUBSCRIPT = divide start_ARG italic_η end_ARG start_ARG 2 end_ARG ( italic_τ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) and Tπ2=η2(τx+τy)subscript𝑇𝜋2𝜂2subscript𝜏𝑥subscript𝜏𝑦T_{\frac{\pi}{2}}=\frac{\eta}{2}(\tau_{x}+\tau_{y})italic_T start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT = divide start_ARG italic_η end_ARG start_ARG 2 end_ARG ( italic_τ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) operators acting on the valley isospin subspace, with τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT being valley Pauli matrices, and with some parameter η𝜂\etaitalic_η. Since these two rotation generators do not commute, [Tπ6,Tπ2]=η2iτz0T_{\frac{\pi}{6}},T_{\frac{\pi}{2}}]=\eta^{2}i\tau_{z}\neq 0italic_T start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 6 end_ARG end_POSTSUBSCRIPT , italic_T start_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT ] = italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_i italic_τ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≠ 0, we suppose that under proper arrangement of the junction, a nonzero geometric phase may emerge in our system, leading to effective valley manipulation.

To have the possibility non-commutative valley rotations and to make our analysis more realistic we planned another device presented in Fig. 10.

Refer to caption
Figure 10: Setup designed for testing a possibility of valley qubit manipulation via geometric phase accumulation. It is based on corner-shaped junction (a) with both terminations of armchair type, while the phase accumulation occurs when circulating on a closed loop (d) near the junction.

It contains a rhombus-shaped flake with crystal lattice oriented in such a way (rotated by π6𝜋6\frac{\pi}{6}divide start_ARG italic_π end_ARG start_ARG 6 end_ARG with respect to the original lattice) that two armchair terminations can be arranged along θ=π3𝜃𝜋3\theta=\frac{\pi}{3}italic_θ = divide start_ARG italic_π end_ARG start_ARG 3 end_ARG, and 2π32𝜋3\frac{2\pi}{3}divide start_ARG 2 italic_π end_ARG start_ARG 3 end_ARG. Moreover, a lateral junction is formed along the crossing of such armchair terminations leading to a corner-shaped junction (with both sides of armchair type) as presented in Fig. 10(a). The corner junction is modelled in a similar manner as the original device with linear dipoles along the both armchair sides.

To enable us to work in a valley degenerate subspace, the external magnetic field B=2.4𝐵2.4B=2.4italic_B = 2.4 T is applied to get appropriate level crossing (spin-down subspace: |K|K\!\downarrow\rangle| italic_K ↓ ⟩ and |K|K^{\prime}\!\downarrow\rangle| italic_K start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↓ ⟩ as presented by orange line in Fig. 10(b). These levels were calculated for an electron localized in position A – see Fig. 10(c). Moving the electron to other positions and therefore controlling its position is realized by applying control voltages to four device gates. This way we can adiabatically move the electron to positions B, C, and D, and after many cycles the electron path resembles a closed loop (orange curve) presented in Fig. 10(d). In position A, the electron is farthest from the junction, while in positions B and D it is closest to the sides of the corner junction. When we calculate the energy levels as a function of the magnetic field for an electron in the D position, we may observe the influence of the nearby terminations manifested by the mixing (anticrossing) of energy levels – blue line in Fig. 10(b). Matrix elements calculated along the electron path presented in Fig. 10(e) show that forward (ABC path) and backward (CDA path) moves are represented by non commutative operators; e.g., at B position Tforwardη(12σx+σy)similar-to-or-equalssubscript𝑇forwardsuperscript𝜂12subscript𝜎𝑥subscript𝜎𝑦T_{\mathrm{forward}}\simeq\eta^{\prime}(\frac{1}{2}\sigma_{x}+\sigma_{y})italic_T start_POSTSUBSCRIPT roman_forward end_POSTSUBSCRIPT ≃ italic_η start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) and Tbackwardη(σx+12σy)similar-to-or-equalssubscript𝑇backwardsuperscript𝜂subscript𝜎𝑥12subscript𝜎𝑦T_{\mathrm{backward}}\simeq-\eta^{\prime}(\sigma_{x}+\frac{1}{2}\sigma_{y})italic_T start_POSTSUBSCRIPT roman_backward end_POSTSUBSCRIPT ≃ - italic_η start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ).

Refer to caption
Figure 11: Valley isospin evolution when traveling along the closed path that enables accumulating of non-Abelian Berry phase (orange curve). When return path is the same, without circling the loop, the valley changes are much smaller (blue).

Now, let us verify if passing next to one side of a corner junction and then next to another induces effective valley isospin rotations; i.e., whether returning to initial position (ABCDA loop) results in the same or a rotated valley. Presented in Fig 11 is the evolution of valley isospin of an electron confined in the nanodevice from Fig. 10. After passing ABCDA loop (blue curve) multiple times, one observes effective valley manipulation which gradually changes after the completion of each loop. For comparison we also calculated valley evolution in the case of a return along the same path ABCBA (without making a loop – see blue curve in Fig. 10(d)). In this case we still observe some valley changes (blue curve in Fig. 11) but with smaller amplitudes.

VI Summary

In the above work, we have studied a single-electron system in a TMDC lateral heterostructure, gate-defined quantum dot from the point of view of a valley-qubit implementation. Utilizing the time-dependent Schrödinger equation in a tight-binding approximation coupled with the Poisson equation (which models a realistic dielectric environment), we were able to describe the proposed nanodevice with an in-plane TMDC heterojunction, four-gate QD geometry, and time-modulated electric potential. We have shown that when the built-in dipole moment at the junction is oriented along the armchair interface, oscillatory pumping of the electron density into the junction area induces intervalley transitions. These transitions are resonant in nature and lead to Rabi oscillations with transition period dependent on electron strength overlapping with the junction area. Lateral interfaces between two different TMDC monolayers have been realized experimentally. However, the real interfaces are not always straight and frequently composed of shorter (10similar-toabsent10\sim 10∼ 10 nm-long) sections of armchair and zigzag type [23]. Thus, we also verified that intervalley coupling is possible even if the electron interacts only with a fragment (e.g. 5similar-toabsent5\sim 5∼ 5 nm-long) of armchair geometry (leaving the rest of such ”kinked” junction in a zigzag form). We also carefully analyzed what factors are necessary to achieve intervalley transitions. This analysis also led us to suggest another possibility of performing operations on the valley isospin. By properly designing a corner junction and carefully manipulating the electron to move along a closed loop near both sides of the junction, it is possible to accumulate the non-Abelian geometric phase and thus effectively rotate the valley.

Currently, another possibility to apply our scheme requiring sharp electronic potential modulations has emerged. Twisted TMDC bilayers seems to be a promising platform to study intervalley physics [53]. In moiré materials K𝐾Kitalic_K valleys are further splitted into two K±limit-from𝐾plus-or-minusK\pmitalic_K ± points (forming so-called moiré BZ) that lie much closer to each other in k𝑘kitalic_k-space and are possibly easier to couple. Moreover, the moiré superlattice, that emerges in real space, may also play a role in inducing intervalley transitions. The benefit of this approach could lie in its compatibility with mechanically exfoliated TMDC crystals rather than epitaxially grown TMDCs, which are known to be of inferior quality.

VII Acknowledgements

The authors would like to thank Oscar Ávalos-Ovando for enlightening discussions. This work has been supported by National Science Centre, Poland, under Grant No. 2021/43/D/ST3/01989. This research was supported in part by PL-Grid Infrastructure.

References